Accepted Manuscript Predictions of water/oil interfacial tension at elevated temperatures and pressures: A molecular dynamics simulation study with biomolecular force fields Konstantinos D. Papavasileiou, Othonas A. Moultos, Ioannis G. Economou PII:
S0378-3812(17)30182-6
DOI:
10.1016/j.fluid.2017.05.004
Reference:
FLUID 11475
To appear in:
Fluid Phase Equilibria
Received Date: 6 April 2017 Revised Date:
29 April 2017
Accepted Date: 4 May 2017
Please cite this article as: K.D. Papavasileiou, O.A. Moultos, I.G. Economou, Predictions of water/oil interfacial tension at elevated temperatures and pressures: A molecular dynamics simulation study with biomolecular force fields, Fluid Phase Equilibria (2017), doi: 10.1016/j.fluid.2017.05.004. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
Predictions of water/oil interfacial tension at
2
elevated temperatures and pressures: A molecular
3
dynamics simulation study with biomolecular force
4
fields
5
Konstantinos D. Papavasileiou,1 Othonas A. Moultos,2 and Ioannis G. Economou1,3,*
M AN U
SC
RI PT
1
6
1
7
Nanotechnology, Molecular Thermodynamics and Modelling of Materials Laboratory,
8
GR-15310 Aghia Paraskevi Attikis, Greece
9
2
TE D
National Center for Scientific Research “Demokritos”, Institute of Nanoscience and
Engineering Thermodynamics, Process & Energy Department, Faculty of Mechanical,
Maritime and Materials Engineering, Delft University of Technology, Leeghwaterstraat
11
39, 2628CB Delft, The Netherlands
12
3
14 15
Texas A&M University at Qatar, Chemical Engineering Program, Education City, PO
AC C
13
EP
10
Box 23874, Doha, Qatar
16 17
*To whom correspondence should be addressed
18
IGE:
[email protected] 1
ACCEPTED MANUSCRIPT
1
Abstract The interfacial properties of water/oil mixtures is a topic of significant interest for the oil and gas
3
and chemical industry, as they are required for performing process calculations. However, the
4
reported data at high temperatures and pressures are scarce. The present study focuses on
5
simulating the interfacial tension (IFT) of mixtures of water with i) toluene (water/toluene), ii) n-
6
dodecane
7
(water/toluene/n-dodecane), at 1.83 MPa and temperatures ranging from 383.15 to 443.15 K.
8
Molecular Dynamics (MD) simulations with atomistic molecular models, developed primarily
9
for biomolecular systems were employed. In the simulations performed, the effects of the water
10
model and the scaling of interatomic interactions by introducing a binary interaction parameter,
11
kij , were assessed for the accurate reproduction of experimental data. The combination of the
12
TIP4P/2005, SPC/E and TIP3P water models with the GAFF and LIPID14 force fields for
13
toluene and n-dodecane respectively, coupled with appropriate binary interaction parameters, kij ,
14
for the interaction of carbon with oxygen atoms were found to yield accurate results in the case
15
of binary mixtures. For the water/toluene/n-dodecane mixture, all force field combinations tested
16
resulted in overestimated IFT values for the whole range of state points examined. Our
17
simulations show that these widely used force fields, originating from the world of biomolecular
18
simulations, are suitable candidates in the study of binary water/oil mixtures. Nevertheless, the
19
introduction of the kij scaling parameter is not sufficient to allow the accurate reproduction of
20
experimental IFT data for ternary water/oil/oil mixtures.
21
Keywords: Interfacial tension; toluene; n-dodecane; Molecular dynamics simulations
and
iii)
a
50:50
%
wt
toluene:n-dodecane
mixture
AC C
EP
TE D
M AN U
SC
(water/n-dodecane)
RI PT
2
22
2
ACCEPTED MANUSCRIPT
1.
2
The 9th Industrial Fluid Properties Simulation Collective Challenge [1] aimed at testing the
3
ability of molecular modeling approaches to predict water/oil interfacial tension (IFT) at elevated
4
temperatures and pressures for three different oils/water mixtures, namely water with toluene, n-
5
dodecane and a 50:50 % wt mixture of n-dodecane and toluene. The IFT, γ, of water/oil systems
6
is one of the basic physical properties required for performing process design calculations in oil
7
and gas and chemical industry [2]. IFT is defined as the force per unit length that is required to
8
increase the surface area between two immiscible fluids in contact [3]. It is an essential property
9
for the description of liquid-liquid interfaces, as it controls contact phase hydrodynamics and
10
mass transfer [4]. Experimentally, IFT is measured by a variety of methods, such as the spinning
11
drop, the du Noüy ring, the Wilhelmy plate etc. [5, 6]. Previous IFT studies on binary water/n-
12
alkane [7-9], water/toluene [8, 10, 11] and ternary mixtures of water/n-alkane/toluene [12]
13
referred to ambient conditions. However, at high temperatures and pressures only a few data
14
have been reported [4, 7, 13-18]. Typically, these conditions involve temperatures that reach and
15
exceed 370 K, while pressures applied are normally beyond 0.1 MPa in order to maintain the
16
liquid state. A common thread in these studies is the IFT tendency to decrease with increasing
17
temperature at constant pressure. Pressure effects are shown to be less pronounced [7], with IFT
18
increasing with increasing pressure at constant temperature [7, 19-21].
SC
M AN U
TE D
EP
AC C
19
Introduction
RI PT
1
Several simulation studies have been dedicated to the examination of the IFT water/oil
20
mixtures. These include application of Monte Carlo (MC) [22, 23], Molecular Dynamics (MD)
21
[24-28] and mesoscale modelling techniques such as Dissipative Particle Dynamics (DPD) [29]
22
and Coarse-Grained Molecular Dynamics (CGMD) [30, 31]. CGMD methods have been
23
enriched recently with force fields developed with the SAFT-γ Mie equation of state (eos) top3
ACCEPTED MANUSCRIPT
down approach, which has been successfully applied in MD simulations for the prediction of
2
several water/oil mixture properties [32]. Within the framework of this challenge, we calculated
3
the IFT of these mixtures by means of interfacial MD simulations, a robust and reliable
4
simulation technique in the study of water/oil systems.
RI PT
1
The accuracy of the molecular models, describing the intra- and inter-molecular
6
interactions, employed in simulations are of outmost importance for attaining accurate IFT
7
values [33]. To this extend, our attempts focused on incorporating and benchmarking the
8
performance of well-established atomistic molecular models coming from the world of
9
biomolecular simulations in systems of interest to the chemical/petrochemical industry. More
10
specifically, for the description of the organic phase we employed the General AMBER force
11
field (GAFF) [34] and the Lipid14 [35] force field for toluene and n-dodecane, respectively.
12
GAFF was developed for the incorporation of small organic molecules to existing AMBER force
13
fields for proteins and nucleic acids. In early attempts of lipid simulations for a range of
14
phospholipid bilayers GAFF showed great potential in the reproduction of experimental
15
membrane properties, such as the area per lipid, the area compressibility modulus, lipid order
16
parameters and others [36, 37]. However, reoccurring issues at long timescales and constant
17
pressure such as a liquid-to-gel phase transition above the phase transition temperatures were
18
observed [38-40]. Similar issues were observed also for the OPLS-AA force field for n-alkanes
19
longer than n-hexane, thus prompting for the development of the L-OPLS force field [41], which
20
was parameterized in order to accurate reproduce liquid densities and viscosities of long
21
hydrocarbons [42]. Therefore, dedicated efforts to tackle these problems led to the development
22
of a new family of lipid force fields based on GAFF, with Lipid14 being its latest variant.
23
Lipid14 features optimized alkyl chain dihedral angle terms and Lennard-Jones (LJ) parameters,
AC C
EP
TE D
M AN U
SC
5
4
ACCEPTED MANUSCRIPT
allowing for tensionless simulations of lipid systems. The key features of these models are their
2
enhanced compatibility with and transferability to a wide range of physical systems, ranging
3
from proteins, lipids to organic compounds [34, 35, 43].
RI PT
1
No prior attempts to utilize the Lipid14 in calculations of non-biological systems have
5
been reported in literature. Recently, Kumar et al. [44] evaluated GAFF’s performance in the
6
prediction of liquid−vapor saturation properties of naphthalene derivatives. GAFF exhibited high
7
accuracy of the liquid−vapor saturation properties and thus was proved to be a reasonable choice
8
for describing the phase behavior of systems relevant to the oil and gas industry. In addition, the
9
computational benchmarking of 146 organic liquids by Caleman et al. [45] showed that GAFF
10
performs well in the prediction of several properties such as density, enthalpy of vaporization,
11
heat capacities, and others.
M AN U
SC
4
We therefore grasped the opportunity of this 9th Challenge in order to evaluate the
13
performance of these models in systems of industrial interest, which briefly encapsulates the
14
novelty of the present work. In addition, we have examined the effect of the non-bonded LJ
15
interactions long-range treatment, the simulation duration and the impact of the choice of the
16
water model on the calculated IFTs.
17
2. Simulation methods
EP
AC C
18
TE D
12
2.1. Force fields
19
Toluene structure was initially optimized in the gas phase at the B3LYP/6-31G* level of theory
20
[46-48], by means of the Gaussian 09 [49] suite of programs. Then, atomic charges were
21
obtained from the optimized geometry at the HF/6-31G* level of theory according to the
22
Mertz-Kollman population analysis scheme [50, 51]. Consecutively, partial atomic charges were 5
ACCEPTED MANUSCRIPT
derived according to the RESP protocol [52], utilizing the ANTECHAMBER module [53] of the
2
AMBER12 suite of programs [54]. GAFF parameters [34] were then assigned to toluene, the
3
atom types of which are presented in Table S1 of the Supporting Information. Topologies were
4
generated by means of the tLEaP module [54] and are illustrated in Table S2, along with the
5
calculated partial charges.
RI PT
1
n-dodecane was described by means of the Lipid14 force field parameters [35]. Partial
7
charges for the n-dodecane chain were generated according to the original derivation procedure
8
[35]. Specifically, fifty randomly selected n-dodecane molecules were extracted from the initial
9
cubic box conformation and were subsequently used for the charge calculation. The Electrostatic
10
Potential (ESP) [50, 51] was calculated directly from each structure at the HF/6-31G* level of
11
theory without optimization, and partial charges derived using the RESP fitting procedure [52]
12
for each chain. Charges were then taken as an average over all fifty RESP fits. The structure and
13
partial charges of n-dodecane are illustrated in Table S3 of the Supporting Information.
TE D
M AN U
SC
6
For the representation of water molecules the TIP4P/2005 [55], TIP3P [56] and SPC/E
15
[57] models were employed. Our starting point was the TIP4P/2005 force field, which is a four-
16
site rigid model with a LJ sphere fixed on the oxygen atom. The electrostatic contributions are
17
implemented by positive partial charges assigned to the hydrogens and a negative partial charge
18
on the additional site, located on the bisector of the H-O-H angle at a distance 0.1546 Å from the
19
oxygen atom. The TIP4P/2005 model has been extensively tested and found to be very accurate
20
for several water properties, including the vapor-liquid IFT [58-60]. On the other hand, the
21
SPC/E and TIP3P force fields assume a single LJ site on the oxygen atom and three charges on
22
oxygen and hydrogen atoms, respectively, while bond lengths and bond angles are kept fixed.
23
The SPC/E force field, likewise to TIP4P/2005, is considered one of the best, two-body, non-
AC C
EP
14
6
ACCEPTED MANUSCRIPT
polarizable, water models [61] and has also proved to be accurate in capturing the vapor-liquid
2
IFT [59]. SPC/E does not use the experimental geometry of the water but instead two simple
3
values for the bond length (1 Å) and angle (109.47˚). The TIP3P is the most commonly used
4
water model in biomolecular simulations [61] and it was primarily developed to reproduce its
5
density and the vaporization enthalpy at ambient temperature and pressure conditions [56, 61].
6
However, several studies have highlighted the TIP3P’s poor performance in describing important
7
water properties, thereby expressing serious concerns about its more generalized use [58, 61].
8
Despite its known deficiencies, given that the Lipid14 and GAFF force fields have been
9
parameterized for use primarily with the TIP3P water model [43] and that this combination is
10
utilized in aqueous membrane bilayer simulations [36], this water model was also included in
11
this study. The force field parameters of all the species involved are collected and presented in
12
Tables S4 and S5 of the Supporting Information document.
M AN U
SC
RI PT
1
Standard Lorentz–Berthelot combining rules [62] were used for the well depth ε and the
14
size parameter σ to describe non-bonded LJ interactions between sites of different type i and j
15
according to the expressions:
ε ij = ε iiε jj and σ ij =
EP
16
TE D
13
σ ii + σ jj 2
.
(1)
The Berthelot (geometric mean) combining rule has been shown to work well for
18
various properties of several binary mixtures of components, having atoms of similar sizes and
19
ionization potentials [63-69]. However, it has been proven to be insufficient in the description of
20
dispersion interactions between polar and non-polar components [63]. Previous simulation
21
studies have shown that the combination of TraPPE n-alkanes and SPC/E water mixtures yields
22
water solubilities that are underestimated by approximately an order of magnitude compared to
23
their experimental values [70]. The source of this behavior may be primarily attributed to the
AC C
17
7
ACCEPTED MANUSCRIPT
inaccurate force field description of electrostatic and induction effects of water in n-alkane [71,
2
72], as well as the crudely modeled water – n-alkane interactions based on the Lorentz –
3
Berthelot combining rules. In order to attain better agreement with experiment, it has become
4
customary to introduce corrections in terms of the deviation from the Berthelot rule [70],
5
expressed as:
RI PT
1
6
ε ij = kij ε iiε jj
7
where, kij is a binary interaction parameter fitted to the experimental data. Such a parameter was
8
also used in the present study, as it was previously found that the interfacial properties of the
9
water/n-decane mixture were not sensitive to size parameter σ and the standard Lorentz
11
SC
M AN U
10
(2)
combining rule was employed for the description of unlike sites [26]. 2.2. Simulation details
The initial conformations of all systems were constructed by means of the Amorphous and
13
Interface Builder modules integrated in Scienomics MAPS software package [73], according to
14
which the constituent molecules were first inserted into the cubic boxes using a modified
15
configuration bias scheme [74], which were later merged to produce final orthorhombic cells of
16
the interfaces phases. The number of molecules used was 2000 for water and 250 for toluene and
17
n-dodecane, respectively. In the ternary water/toluene/n-dodecane mixture, 162 toluene and 88 n-
18
dodecane molecules were used, respectively, in order to properly account for the 50:50 (wt/wt)
19
toluene/n-dodecane blend requirement of the challenge. Then, the generated GAFF, Lipid14 and
20
TIP4P/2005, TIP3P and SPC/E topologies and coordinate files were converted to the
21
GROMACS format by means of the ACPYPE utility [75] (Table S5). All topology and
22
coordinate files are available as a Mendeley dataset [76].
AC C
EP
TE D
12
8
ACCEPTED MANUSCRIPT
1
The liquid-liquid IFT was obtained by interfacial MD simulations using GROMACS
2
5.0.7 software [77-81]. The simulations were performed in orthorhombic boxes, with periodic
3
boundary conditions imposed in all directions (Figure 1). Prior to production runs and in order to eliminate any close contacts between atoms, all
5
systems were subjected to steepest descent energy minimization for 20,000 steps. Then, short
6
equilibration simulations in the canonical (NVT) and isobaric-isothermal-isointerface area
7
(NPNAT) ensembles were performed, where PN is the input normal pressure and A the interfacial
8
area. Specifically, all systems were gradually heated to the target temperatures for 100 ps in the
9
NVT ensemble using the Berendsen thermostat [82], with the coupling constant set to 1 ps. LJ
10
interactions were cut-off at 14 Å and the Particle Mesh Ewald (PME) method [83, 84] was
11
applied for the calculation of long-range electrostatic interactions at the same cut-off distance.
12
The LINCS algorithm [85] restrained all bonds involving hydrogen atoms. The integration step
13
of all simulations was set to 2 fs.
TE D
M AN U
SC
RI PT
4
All systems were then equilibrated for a period of 5 ns in the NPNAT ensemble, at
15
constant pressure of 1.83 MPa using a Berendsen barostat [82] with the coupling constant set to
16
1 ps. Pressure coupling was isotropic in the x and y direction, but semi-isotropic in the z
17
direction, which is perpendicular to the liquid-liquid interface. During this period, the density of
18
both liquid phases converged to a mean value, along with the energy of the system (Figure 1).
20
21
AC C
19
EP
14
The IFT, γ , was evaluated by using the diagonal stress tensor elements ( Pxx , Pyy , and
Pzz ) through the following expression [86-88]:
γ=
Pxx + Pyy 1 Lz Pzz − 2 2
(3)
9
ACCEPTED MANUSCRIPT
where, Lz is the length of the simulation box in the z direction, normal to the water/oil interface
2
which forms along the xy plane. The factor ½ is due to the presence of two interfaces and angle
3
brackets represent the ensemble average.
RI PT
1
For the purposes of the challenge, IFTs were calculated without employing any
5
modifications to LJ combining rules from a single 5 ns run per state point, in the NPNAT
6
ensemble using the Nosé–Hoover thermostat [89], while maintaining the semi-isotropic pressure
7
coupling with a Berendsen barostat [82]. IFT values and statistical uncertainties (95% confidence
8
intervals) were obtained by dividing the production period of the simulations into five blocks
9
(Table S6 of Supporting Information). After the submission to the challenge, longer production
10
runs of 10 ns were performed. Each run was executed on 100 Intel Xeon E5-2680v2 cores and
11
required approximately two wall-clock hours for completion. The simulation length was found to
12
have no impact on IFT results (Table S7 of Supporting Information). It is important to note here,
13
that the kij binary interaction parameter was fitted after the end of the challenge and upon
14
disclosure of the experimental data.
TE D
M AN U
SC
4
The cut-off of the LJ interactions has a minor impact on the molecular arrangement, but
16
significantly influences the computed values of properties such as the IFT. This is more profound
17
in simulations of inhomogeneous systems, leading to incorrect estimation of interfacial
18
properties [59]. Therefore, instead of using analytical tail corrections, we applied the Particle-
19
mesh Ewald summation for the attractive LJ interactions (LJ-PME) [90]. This technique is
20
computationally robust and has been successfully employed in the calculation of properties such
21
as IFT [91]. However, previous works have shown that using LJ-PME with the GAFF force
22
field, which has been optimized for a cut-off approach, leads to overestimated IFT for several
23
organic liquids [92]. We investigated the effect of using the LJ-PME over the cut-off to the long-
AC C
EP
15
10
ACCEPTED MANUSCRIPT
range LJ potential, using a radius of 16 Å a practice that has shown to be equally efficient [90].
2
Our IFT results showed no dependence on the long-range LJ treatment chosen (Tables S7 of
3
Supporting Information). Snapshots of the equilibrated structures of all systems are illustrated in
4
Figure 1.
5
RI PT
1
3. Results and Discussion
In this section, the MD simulations results for the series of water/oil mixtures investigated are
7
presented. All IFT calculations are at the reference temperatures of 383.15, 403.15, 423.15 and
8
443.15 K and pressure of 1.83 MPa. 3.1.
M AN U
9
SC
6
Berthelot combining rule
The density profiles illustrated in Figure 1, show that the bulk region densities of water
11
(TIP4P/2005) and n-dodecane (Lipid14) are close to their experimental values [93]. Toluene
12
density is slightly underestimated by GAFF compared to its experimental value, a finding also
13
reported by Caleman et al. at 298.15 K and 1 bar [45]. Furthermore, the density distribution of
14
the water/toluene/n-dodecane mixture components is in agreement with previous simulation
15
studies of water/oil interfaces, particularly with regard to the tendency of the toluene molecules
16
to accumulate at the water interface (Figure 1c) [28]. This accumulation phenomenon at water/oil
17
interfaces is attributed to the weak hydrogen bonding between water and toluene at the interface
18
[49]. These density distribution profiles are essentially the same for SPC/E, with the partial
19
density being slightly underestimated compared to the experimental values at each temperature
20
(Figure S3 of Supporting Information). On the other hand, in the TIP3P water systems the
21
underestimation of water density becomes more significant, a feature which should have an
22
important effect on the accurate calculation of IFTs (Figure S4).
AC C
EP
TE D
10
11
ACCEPTED MANUSCRIPT
From the results presented in Table 1, it is indicated that the IFT values in all systems
2
and for all force fields studied decrease as temperature increases, reproducing the experimentally
3
observed trends. Figure 2 shows that there is an almost linear dependence between γ and T; this
4
behavior is in agreement with previous experimental studies on water – n-alkane mixtures [13].
5
The water/n-dodecane mixture exhibits higher IFT compared to water/toluene, with the
6
water/toluene/n-dodecane mixture adopting intermediate values.
RI PT
1
These observations are in line with the experimental values used as a benchmark for the
8
purposes of the challenge (Figure 2). However, even though the combination of the GAFF and
9
Lipid14 AMBER force fields with TIP4P/2005 appears to capture the correct trend, absolute IFT
10
values are overestimated compared to the experimental ones by 20%, 27% and 30% on average
11
for the water/toluene, water/n-dodecane and water/toluene/n-dodecane mixtures, respectively.
12
The SPC/E displays the same behavior with TIP4P/2005. Furthermore, the overestimation is
13
increasing with increasing temperature. It is also clear that the TIP3P water model has a
14
profound impact on IFT. Specifically, the IFT results of all mixtures are now underestimated but
15
are slightly closer in absolute value compared to the experimental benchmark data. Furthermore,
16
the TIP3P model leads to an improvement in the description of the IFT temperature dependence,
17
especially in the case of the water/toluene mixture.
M AN U
TE D
EP
It is useful to note here that all water models have similar σ values for water oxygen,
AC C
18
SC
7
19
with differences in their ε values (Table S4 of the Supporting Information). In fact, the ε values
20
decrease in the order εTIP4P/2005>εSPC/E>εTIP3P. Thus, by altering the εij of the toluene and n-
21
dodecane carbon and water oxygen atoms, respectively, one can fine-tune these models in order
22
to achieve better agreement with the experimental IFT values.
23
12
ACCEPTED MANUSCRIPT
1
3.2.
Modified Berthelot combining rule
The simulations with scaled binary interaction parameters of the toluene and n-dodecane carbon
3
and water oxygen atoms, respectively, resulted in better agreement with experiment. The kij
4
values led to the most accurate results are presented in Table 2. One can observe that the
5
TIP4P/2005 and SPC/E models were found to perform better with different kij for the aromatic
6
and alkyl carbons with water oxygen. On the other hand, mixtures with the TIP3P model were in
7
better agreement by setting a uniform kij = 0.8 value for all carbon atoms with water oxygen
8
interactions. From the results listed in Table 3 and depicted in Figure 3 become clear that the
9
fine-tuning of the kij parameter can produce much improved water/oil IFT values. Nevertheless,
10
upon closer examination, one can isolate subtle differences introduced to the simulations by the
11
choice of water model.
M AN U
SC
RI PT
2
The average absolute difference of simulated and experimental IFT values is ~2% for
13
the water/toluene and water/n-dodecane mixtures for all water models used. This translates to an
14
absolute difference of 1 mN/m. Despite the overall improvement, there seems to be a
15
discrepancy that is not tackled even after the introduction of the kij parameter. More specifically,
16
it is observed that the simulations including TIP4P/2005 and SPC/E systems show better
17
agreement with experiments at lower temperatures, which is not followed at higher temperatures.
18
This results in a different temperature dependence of IFT, which diverges from the experimental
19
trend. The TIP3P model coupled with GAFF and Lipid14 for toluene and n-dodecane,
20
respectively, results in better reproduction of the data for the entire temperature range examined.
21
This can be attributed to the compatibility of the TIP3P model with the AMBER family of force
22
fields.
AC C
EP
TE D
12
13
ACCEPTED MANUSCRIPT
Deviation from experiment becomes more pronounced in the case of the
2
water/toluene/n-dodecane mixture. Calculated IFT values deviate by about 8%, which translates
3
to approximately 3 mN/m on average for all systems. Similarly to the findings for the binary
4
mixtures, the TIP4P/2005 / GAFF / Lipid14 and SPC/E /GAFF / Lipid14 combinations result to
5
IFTs that are closer to the experimental values at lower temperatures, while adopting slightly
6
higher values compared to experiment as temperature increases (Table 3 and Figure 3). The
7
TIP3P / GAFF / Lipid14 combination shows the most consistent behavior, but IFT values are
8
still overestimated by approximately 6%. This indicates that even though the simple kij scaling
9
approach is adequate for improving results for the binary water/oil mixtures, it is not enough to
SC
M AN U
10
RI PT
1
capture the more complex intermolecular interactions in ternary water/oil/oil mixtures. There appear to be significant differences in the interfacial loading of toluene depending
12
on the water model used, which explains the differences in IFT. Given the different electrostatics
13
of each water model, the attractive “weak hydrogen bonding” between the aromatic toluene and
14
water molecules [28] is either enhanced (TIP4P/2005, SPC/E) or reduced (TIP3P). This explains
15
the difference in toluene density observed with either water model (Figure 4a). The preference of
16
toluene to accumulate near the water interface in the water/toluene/n-dodecane mixture, not only
17
explains the intermediate IFT values of the ternary compared to the binary mixtures, but also
18
renders toluene loading at the interface as a controlling factor of the IFT.
EP
AC C
19
TE D
11
From our simulations, it becomes clear that the use of the kij adjustable parameter alone
20
is insufficient to predict accurately the experimental IFT data for ternary water/oil/oil mixtures.
21
Fine-tuning the kij parameter results into a toluene density profile at the water/oil interface that
22
is now common for all systems (Figure 4b). This highlights the role of toluene as an active
23
interfacial agent in the ternary mixture. 14
ACCEPTED MANUSCRIPT
1
4. Conclusions The present work entails extensive MD simulations of representative water/oil binary and ternary
3
mixtures for the calculation of IFT. Our efforts concentrated primarily in incorporating atomistic
4
force fields that have been primarily developed for biomolecular simulations in addressing
5
industrially related properties like the IFT. The GAFF and Lipid14 force fields were employed
6
for the description of toluene and n-dodecane, respectively. The partial density profiles of the oil
7
phases were found to be in agreement with the respective experimental data. These molecular
8
models were combined with three fixed-point-charge water models, namely the TIP4P/2005,
9
TIP3P and SPC/E. The TIP3P model does not reproduce the experimental water density as
10
accurately as its TIP4P/2005 and SPC/E counterparts do. Water model parameters are crucial in
11
the prediction of IFT properties, as they are responsible for the significant overestimation
12
(TIP4P/2005 and SPC/E) or the moderate underestimation (TIP3P) of IFT values compared to
13
the experimental data.
TE D
M AN U
SC
RI PT
2
Modifying the cross LJ interactions by appropriately adjusting the binary interaction
15
parameter is a simple but effective way to improve the accuracy of the IFT predictions for binary
16
mixtures. However, even by following this approach, the use of the TIP4P/2005 and SPC/E
17
models leads to slightly overestimated IFT values at 443.15 K, weakly diverging from the
18
experimental weak temperature IFT dependence. Surprisingly, the combination of the less
19
accurate TIP3P water model with the aforementioned AMBER force fields, using a uniform
20
kij = 0.8 parameter of the scaling of carbon-water oxygen interactions captures the overall IFT
21
behavior more accurately compared to the other two water models. TIP3P’s deficiency in
22
accurately reproducing water density at the conditions examined is compensated by its
AC C
EP
14
15
ACCEPTED MANUSCRIPT
1
compatibility with the AMBER force fields and aids towards calculation of IFTs that are in
2
better agreement with experiment. As far as the water/toluene/n-dodecane mixture is concerned, all simulations, regardless
4
of the kij scaling used, led to overestimated IFT values compared to experimental findings. The
5
TIP4P/2005 and SPC/E model behavior in higher temperatures pertains, while the TIP3P model
6
shows better agreement. The overestimation of IFT for the ternary mixture cannot be
7
compensated by the introduction of a simple kij binary parameter. The source of this discrepancy
8
can be attributed to the initial atomistic force field parametrization of the pure components and to
9
the actual combining rules used. Therefore, improved models that include IFT data into their
SC
M AN U
10
RI PT
3
initial setup may lead to more accurate mixture calculations.
11 12
AUTHOR INFORMATION Corresponding Authors
14
*IGE:
[email protected]
15
Author Contributions
16
The manuscript was written through contributions of all authors. All authors have given approval
17
to the final version of the manuscript.
19
EP
AC C
18
TE D
13
Acknowledgments
20
This work was supported by computational time granted from the Greek Research & Technology
21
Network (GRNET) in the National High Performance Computing facility - ARIS - under project
22
ID002043. We are also thankful to the High Performance Computing center of Texas A&M
23
University at Qatar for generous resource allocation. Scienomics SARL is acknowledged for
24
providing MAPS software used to generate initial structures of the systems simulated.
25 16
ACCEPTED MANUSCRIPT
1
Table 1. Calculated IFT values for the various mixtures with the three different water models
2
examined, without any scaling of the intermolecular interaction parameters ( kij =1).
γ (mN/m) water/toluene
20.8 ± 0.6
443.15
25.9 ± 0.7
18.8 ± 0.6
SPC/E
29.0 ± 0.6
29.4 ± 0.8
15.3 ± 0.6
27.1 ± 0.7
Experiment
31.2 ± 1.0
31.1 ± 0.5
SC
22.8 ± 0.3
M AN U
423.15
33.1 ± 0.7
TIP3P
403.15
TIP4P/2005
383.15
RI PT
T (K)
28.6 26.4 23.8
24.8 ± 0.9
20.2
36.8 ± 1.2
45.3 ± 1.2
40.0
33.4 ± 0.7
42.6 ± 0.7
443.15
32.9 28.5
31.4
water/toluene/n-dodecane 28.9 ± 0.6
37.5 ± 0.3
36.4 ± 0.3
26.6 ± 1.1
35.3 ± 0.3
31.6 ± 0.9
36.5
36.1 ± 0.9
38.7 ± 0.8
34.7 ± 0.4
Experiment
SPC/E
26.1 ± 0.4
39.9 ± 1.0
23.6 ± 0.6 20.6 ± 0.4
32.9 ± 0.5 29.0 ± 0.6
Experiment
423.15
29.5 ± 0.6
SPC/E
403.15
TIP4P/2005
AC C
383.15
38.5 ± 1.1
EP
443.15
42.4 ± 0.8
TIP3P
423.15
45.4 ± 0.8
TE D
403.15
47.5 ± 0.8
TIP4P/2005
383.15
TIP3P
water/n-dodecane
29.0 26.1 22.6
3 4
17
ACCEPTED MANUSCRIPT
Table 2. Binary interaction parameter values used for IFT predictions. The atom types refer to
2
the interaction sites of toluene (ca: aromatic, c3: aliphatic) and n-dodecane (cD: aliphatic)
3
carbons with water oxygen (OWT4: TIP4P/2005, OW: TIP3P and SPC/E oxygen types; for the
4
actual force field parameters see Tables S1 and S4 of the Supporting Information).
Water model
SPC/E
Value Pair type Value Pair type Value
5
OWT4-ca
OWT4-cD
1.25
1.05
1.25
OW-c3 0.80 OW-c3
OW-ca
OW-cD
0.80
0.80
OW-ca
OW-cD
1.20
1.05
1.20
AC C
EP
TE D
6
OWT4-c3
SC
TIP3P
Pair type
n-dodecane
M AN U
TIP4P/2005
toluene
RI PT
1
18
ACCEPTED MANUSCRIPT
1
Table 3. Calculated IFT values with 95% confidence intervals using the three different water
2
models, using parameters shown in Table 2.
γ (mN/m) water/toluene
24.6 ± 0.6
443.15
22.7 ± 0.8
26.5 ± 0.4
23.6 ± 0.4
20.2 ± 0.5
24.6 ± 0.7
Experiment
27.2 ± 0.4
SPC/E
26.2 ± 0.3
28.4 ± 0.2
SC
29.6 ± 0.9
M AN U
423.15
28.1 ± 0.9
TIP3P
403.15
TIP4P/2005
383.15
RI PT
T (K)
28.6 26.4 23.8
22.1 ± 0.4
20.2
39.8 ± 1.0
40.0 ± 1.3
40.0
36.2 ± 0.5
36.2 ± 0.7
443.15
31.4 ± 1.1
33.6 ± 0.5
33.2 ± 0.8
30.5 ± 0.8
31.3 ± 0.6
30.8 ± 0.7
27.2 ± 0.8
36.5 32.9 28.5
water/toluene/n-dodecane
32.6 ± 0.6
29.0 ± 0.9
Experiment
SPC/E
28.4 ± 0.4
33.6 ± 0.9
27.4 ± 0.7 23.9 ± 0.6
28.3 ± 0.5 25.9 ± 0.5
31.4
Experiment
423.15
32.7 ± 0.3
SPC/E
403.15
TIP4P/2005
AC C
383.15
31.0 ± 0.8
EP
443.15
33.7 ± 0.9
TIP3P
423.15
36.7 ± 0.8
TE D
403.15
38.7 ± 1.2
TIP4P/2005
383.15
TIP3P
water/n-dodecane
29.0 26.1 22.6
3 4
19
1
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
2
Figure 1. Equilibrated structures and partial densities of all systems studied using the
3
TIP4P/2005 water model, for various temperatures and pressure 1.83 MPa. Water, toluene and n-
4
dodecane molecules are illustrated in red, gray and cyan colors, respectively.
5
20
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
1 2
Figure 2. Calculated IFT of water/toluene, water/n-dodecane and 50:50 (wt/wt) water/toluene/n-
3
dodecane mixtures with the TIP5P/2005, TIP3P and SPC/E water models, without scaling of the
4
intermolecular interaction parameters.
21
AC C
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
1 2
Figure 3. Calculated IFT of water/toluene, water/n-dodecane and 50:50 (wt/wt) water/toluene/n-
3
dodecane mixtures with the TIP5P/2005, TIP3P and SPC/E water models, by scaling the
4
intermolecular carbon – oxygen interaction parameters. Binary interaction parameter values used
5
for IFT predictions depending on the water model are provided in Table 2.
6
22
EP
TE D
M AN U
SC
RI PT
ACCEPTED MANUSCRIPT
Figure 4. Density profiles for the components of the ternary water/toluene/n-dodecane mixture at
3
383.15 K a) without and b) with scaling of the intermolecular carbon – oxygen interaction
4
parameters (Table 2) for all force fields used.
5
AC C
1 2
23
ACCEPTED MANUSCRIPT
TE D
M AN U
SC
RI PT
[1] Industrial Fluid Properties Simulation Collective, http://fluidproperties.org, accessed March 14, 2017. [2] C.Y. Jian, M.R. Poopari, Q.X. Liu, N. Zerpa, H.B. Zeng, T. Tang, Reduction of water/oil interfacial tension by model asphaltenes: the governing role of surface concentration, J. Phys. Chem. B, 120 (2016) 5646-5654. [3] R.J. Hunter, L.R. White, Foundations of colloid science, Oxford University Press, Oxford Oxfordshire, 1987. [4] T. Al-Sahhaf, A. Elkamel, A.S. Ahmed, A.R. Khan, The influence of temperature, pressure, salinity, and surfactant concentration on the interfacial tension of the n-octane-water system, Chem. Eng. Commun., 192 (2005) 667-684. [5] A.W. Adamson, A.P. Gast, Physical chemistry of surfaces, 6th ed., Wiley, New York, 1997. [6] J. Drelich, C. Fang, C.L. White, Measurement of interfacial tension in fluid-fluid systems, in: A. Hubbard (Ed.) Encyclopedia of surface and colloid science, Marcel-Dekker, New York, 2002, pp. 3158-3163. [7] B.Y. Cai, J.T. Yang, T.M. Guo, Interfacial tension of hydrocarbon plus water/brine systems under high pressure, J Chem Eng Data, 41 (1996) 493-496. [8] N.R. Biswal, J.K. Singh, Interfacial behavior of nonionic Tween 20 surfactant at oil-water interfaces in the presence of different types of nanoparticles, RSC Adv., 6 (2016) 113307113314. [9] A. Goebel, K. Lunkenheimer, Interfacial tension of the water/n-alkane interface, Langmuir, 13 (1997) 369-372. [10] N.N. Dutta, G.S. Patil, Effect of phase-transfer catalysts on the interfacial tension of water/toluene system, Can. J. Chem. Eng., 71 (1993) 802-804. [11] J. Saien, S. Akbari, Interfacial tension of toluene + water + sodium dodecyl sulfate from (20 to 50) °C and pH between 4 and 9, J. Chem. Eng. Data, 51 (2006) 1832-1835. [12] S. Gao, K. Moran, Z. Xu, J. Masliyah, Role of bitumen components in stabilizing water-indiluted oil emulsions, Energy Fuels, 23 (2009) 2606-2612. [13] S. Zeppieri, J. Rodríguez, A.L. López de Ramos, Interfacial tension of alkane + water systems, J. Chem. Eng. Data, 46 (2001) 1086-1088. [14] G. Wiegand, E.U. Franck, Interfacial tension between water and non-polar fluids up to 473 K and 2800 bar, Ber. Bunsen-Ges. Phys. Chem., 98 (1994) 809-817. [15] R. Aveyard, D.A. Haydon, Thermodynamic properties of aliphatic hydrocarbon/water interfaces, Trans. Faraday Soc., 61 (1965) 2255-2261. [16] A.S. Michaels, E.A. Hauser, Interfacial tension at elevated pressure and temperature. II. Interfacial properties of hydrocarbon–water systems., J. Phys. Chem., 55 (1951) 408-421. [17] S.S. Susnar, H.A. Hamza, A.W. Neumann, Pressure dependence of interfacial tension of hydrocarbon—water systems using axisymmetric drop shape analysis, Colloids Surf., A, 89 (1994) 169-180. [18] A. Georgiadis, G. Maitland, J.P.M. Trusler, A. Bismarck, Interfacial tension measurements of the (H2O + n-decane + CO2) ternary system at elevated pressures and temperatures, J. Chem. Eng. Data, 56 (2011) 4900-4908. [19] N. Matubayasi, K. Motomura, S. Kaneshina, M. Nakamura, R. Matuura, Effect of pressure on interfacial tension between oil and water, Bull. Chem. Soc. Jpn., 50 (1977) 523-524. [20] F.G. McCaffery, Measurement of interfacial tensions and contact angles at high temperature and pressure, J. Can. Petrol. Technol., 11 (1972) 26-32.
EP
2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46
References
AC C
1
24
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
[21] H.Y. Jennings, The effect of temperature and pressure on the interfacial tension of benzenewater and normal decane-water, J. Colloid Interface Sci., 24 (1967) 323-329. [22] P. Linse, Monte Carlo simulation of liquid–liquid benzene–water interface, J. Chem. Phys., 86 (1987) 4177-4187. [23] F. Biscay, A. Ghoufi, V. Lachet, P. Malfreyt, Monte Carlo calculation of the methane-water interfacial tension at high pressures, J. Chem. Phys., 131 (2009) 124707. [24] J.P. Nicolas, N.R. de Souza, Molecular dynamics study of the n-hexane–water interface: towards a better understanding of the liquid–liquid interfacial broadening J. Chem. Phys., 120 (2004) 2464-2469. [25] S.S. Jang, S.T. Lin, P.K. Maiti, M. Blanco, W.A. Goddard, P. Shuler, Y.C. Tang, Molecular dynamics study of a surfactant-mediated decane−water interface: effect of molecular architecture of alkyl benzene sulfonate, J. Phys. Chem. B, 108 (2004) 12130-12140. [26] A.R. van Buuren, S.J. Marrink, H.J.C. Berendsen, A molecular dynamics study of the decane/water interface, J. Phys. Chem., 97 (1993) 9206-9212. [27] J.L. Rivera, C. McCabe, P.T. Cummings, Molecular simulations of liquid-liquid interfacial properties: water–n-alkane and water-methanol–n-alkane systems, Phys. Rev. E, 67 (2003) 011603. [28] M. Kunieda, K. Nakaoka, Y.F. Liang, C.R. Miranda, A. Ueda, S. Takahashi, H. Okabe, T. Matsuoka, Self-accumulation of aromatics at the oil−water interface through weak hydrogen bonding, J. Am. Chem. Soc., 132 (2010) 18281-18286. [29] H. Rezaei, H. Modarress, Dissipative particle dynamics (DPD) study of hydrocarbon–water interfacial tension (IFT), Chem. Phys. Lett., 620 (2015) 114-122. [30] M. Ndao, J. Devemy, A. Ghoufi, P. Malfreyt, Coarse-Graining the liquid–liquid interfaces with the MARTINI force field: how is the interfacial tension reproduced?, J. Chem. Theory Comput., 11 (2015) 3818-3828. [31] J.C. Neyt, A. Wender, V. Lachet, A. Ghoufi, P. Malfreyt, Quantitative predictions of the interfacial tensions of liquid–liquid interfaces through atomistic and coarse grained models, J. Chem. Theory Comput., 10 (2014) 1887-1899. [32] O. Lobanova, A. Mejía, G. Jackson, E.A. Müller, SAFT-γ force field for the simulation of molecular fluids 6: Binary and ternary mixtures comprising water, carbon dioxide, and nalkanes, J Chem Thermodyn, 93 (2016) 320-336. [33] A. Ghoufi, P. Malfreyt, D.J. Tildesley, Computer modelling of the surface tension of the gas–liquid and liquid–liquid interface, Chem. Soc. Rev., 45 (2016) 1387-1409. [34] J.M. Wang, R.M. Wolf, J.W. Caldwell, P.A. Kollman, D.A. Case, Development and testing of a general Amber force field, J Comput Chem, 25 (2004) 1157-1174. [35] C.J. Dickson, B.D. Madej, Å.A. Skjevik, R.M. Betz, K. Teigen, I.R. Gould, R.C. Walker, Lipid14: the Amber lipid force field, J. Chem. Theory Comput., 10 (2014) 865-879. [36] B. Jójárt, T.A. Martinek, Performance of the general amber force field in modeling aqueous POPC membrane bilayers, J. Comput. Chem., 28 (2007) 2051-2058. [37] S.W.I. Siu, R. Vácha, P. Jungwirth, R.A. Böckmann, Biomolecular simulations of membranes: physical properties from different force fields, J. Chem. Phys., 128 (2008) 125103 [38] S.W. Chiu, S.A. Pandit, H.L. Scott, E. Jakobsson, An improved united atom force field for simulation of mixed lipid bilayers, J. Phys. Chem. B, 113 (2009) 2748-2763. [39] D. Poger, W.F. Van Gunsteren, A.E. Mark, A new force field for simulating phosphatidylcholine bilayers, J. Comput. Chem., 31 (2010) 1117-1125.
AC C
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46
25
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
[40] A.A. Skjevik, B.D. Madej, R.C. Walker, K. Teigen, LIPID11: a modular framework for lipid simulations using Amber, J. Phys. Chem. B, 116 (2012) 11124-11136. [41] S.W.I. Siu, K. Pluhackova, R.A. Bockmann, Optimization of the OPLS-AA force field for long hydrocarbons, J. Chem. Theory Comput., 8 (2012) 1459-1470. [42] O.A. Moultos, I.N. Tsimpanogiannis, A.Z. Panagiotopoulos, J.P.M. Trusler, I.G. Economou, Atomistic molecular dynamics simulations of carbon dioxide diffusivity in n-hexane, ndecane, n-hexadecane, cyclohexane, and squalane, J. Phys. Chem. B, 120 (2016) 1289012900. [43] W.D. Cornell, P. Cieplak, C.I. Bayly, I.R. Gould, K.M. Merz, D.M. Ferguson, D.C. Spellmeyer, T. Fox, J.W. Caldwell, P.A. Kollman, A second generation force field for the simulation of proteins, nucleic acids, and organic molecules, J. Am. Chem. Soc., 118 (1996) 2309-2309. [44] V. Kumar, K.S. Rane, S. Wierzchowski, M. Shaik, J.R. Errington, Evaluation of the performance of GAFF and CGenFF in the prediction of liquid-vapor saturation properties of naphthalene derivatives, Ind. Eng. Chem. Res., 53 (2014) 16072-16081. [45] C. Caleman, P.J. van Maaren, M.Y. Hong, J.S. Hub, L.T. Costa, D. van der Spoel, Force field benchmark of organic liquids: density, enthalpy of vaporization, heat capacities, surface tension, compressibility, expansion coefficient and dielectric constant, J. Chem. Theory Comput., 8 (2012) 61-74. [46] A.D. Becke, Density‐functional thermochemistry. III. The role of exact exchange, J. Chem. Phys., 98 (1993) 5648-5652. [47] P.J. Stephens, F.J. Devlin, C.F. Chabalowski, M.J. Frisch, Ab Initio calculation of vibrational absorption and circular dichroism spectra using Density Functional force fields, J. Phys. Chem., 98 (1994) 11623-11627. [48] W.J. Hehre, R. Ditchfield, J.A. Pople, Self—consistent molecular orbital methods. XII. Further extensions of Gaussian—type basis sets for use in molecular orbital studies of organic molecules, J. Chem. Phys., 56 (1972) 2257-2261. [49] M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G.A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H.P. Hratchian, A.F. Izmaylov, J. Bloino, G. Zheng, J.L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, J.A. Montgomery Jr., J.E. Peralta, F. Ogliaro, M.J. Bearpark, J. Heyd, E.N. Brothers, K.N. Kudin, V.N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari, A.P. Rendell, J.C. Burant, S.S. Iyengar, J. Tomasi, M. Cossi, N. Rega, N.J. Millam, M. Klene, J.E. Knox, J.B. Cross, V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R.E. Stratmann, O. Yazyev, A.J. Austin, R. Cammi, C. Pomelli, J.W. Ochterski, R.L. Martin, K. Morokuma, V.G. Zakrzewski, G.A. Voth, P. Salvador, J.J. Dannenberg, S. Dapprich, A.D. Daniels, Ö. Farkas, J.B. Foresman, J.V. Ortiz, J. Cioslowski, D.J. Fox, Gaussian 09, Gaussian, Inc., Wallingford, CT, USA, 2009. [50] B.H. Besler, K.M. Merz, P.A. Kollman, Atomic charges derived from semiempirical methods, J. Comput. Chem., 11 (1990) 431-439. [51] U.C. Singh, P.A. Kollman, An approach to computing electrostatic charges for molecules, J. Comput. Chem., 5 (1984) 129-145. [52] C.I. Bayly, P. Cieplak, W.D. Cornell, P.A. Kollman, A well-behaved electrostatic potential based method using charge restraints for deriving atomic charges: the RESP model, J. Phys. Chem., 97 (1993) 10269-10280.
AC C
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46
26
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
[53] J.M. Wang, W. Wang, P.A. Kollman, D.A. Case, Automatic atom type and bond type perception in molecular mechanical calculations, J. Mol. Graph. Model., 25 (2006) 247-260. [54] D.A. Case, T.A. Darden, T.E.I. Cheatham, C.L. Simmerling, J. Wang, R.E. Duke, R. Luo, R.C. Walker, W. Zhang, K.M. Merz, B. Roberts, S. Hayik, A. Roitberg, G. Seabra, J. Swails, A.W. Goetz, I. Kolossváry, K.F. Wong, F. Paesani, J. Vanicek, R.M. Wolf, J. Liu, X. Wu, S.R. Brozell, T. Steinbrecher, H. Gohlke, Q. Cai, X. Ye, J. Wang, M.-J. Hsieh, G. Cui, D.R. Roe, D.H. Mathews, M.G. Seetin, R. Salomon-Ferrer, C. Sagui, V. Babin, T. Luchko, S. Gusarov, A. Kovalenko, P.A. Kollman, AMBER12, University of California, San Francisco, 2012. [55] J.L.F. Abascal, C. Vega, A general purpose model for the condensed phases of water: TIP4P/2005, J. Chem. Phys., 123 (2005) 234505. [56] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, M.L. Klein, Comparison of simple potential functions for simulating liquid water, J. Chem. Phys., 79 (1983) 926-935. [57] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, The missing term in effective pair potentials, J. Phys. Chem., 91 (1987) 6269-6271. [58] C. Vega, J.L. Abascal, M.M. Conde, J.L. Aragones, What ice can teach us about water interactions: a critical comparison of the performance of different water models, Faraday Discuss., 141 (2009) 251-276. [59] C. Vega, E. de Miguel, Surface tension of the most popular models of water by using the test-area simulation method, J. Chem. Phys., 126 (2007) 154707. [60] H. Jiang, O.A. Moultos, I.G. Economou, A.Z. Panagiotopoulos, Hydrogen-bonding polarizable intermolecular potential model for water, J. Phys. Chem. B, 120 (2016) 1235812370. [61] C. Vega, J.L.F. Abascal, Simulating water with rigid non-polarizable models: a general perspective, Phys. Chem. Chem. Phys., 13 (2011) 19663-19688. [62] F.T. Smith, Atomic distortion and the combining rule for repulsive potentials, Phys. Rev. A, 5 (1972) 1708-1713. [63] A.J. Haslam, A. Galindo, G. Jackson, Prediction of binary intermolecular potential parameters for use in modelling fluid mixtures, Fluid Phase Equilib., 266 (2008) 105-128. [64] V.K. Michalis, O.A. Moultos, I.N. Tsimpanogiannis, I.G. Economou, Molecular dynamics simulations of the diffusion coefficients of light n-alkanes in water over a wide range of temperature and pressure, Fluid Phase Equilib., 407 (2016) 236-242. [65] O.A. Moultos, I.N. Tsimpanogiannis, A.Z. Panagiotopoulos, I.G. Economou, Atomistic molecular dynamics simulations of CO2 diffusivity in H2O for a wide range of temperatures and pressures, J. Phys. Chem. B, 118 (2014) 5532-5541. [66] O.A. Moultos, G.A. Orozco, I.N. Tsimpanogiannis, A.Z. Panagiotopoulos, I.G. Economou, Atomistic molecular dynamics simulations of H2O diffusivity in liquid and supercritical CO2, Mol. Phys., 113 (2015) 3383-3383. [67] G.A. Orozco, O.A. Moultos, H. Jiang, I.G. Economou, A.Z. Panagiotopoulos, Molecular simulation of thermodynamic and transport properties for the H2O+NaCl system J. Chem. Phys., 141 (2014) 234507. [68] M. Rouha, I. Nezbeda, Second virial coefficients: a route to combining rules?, Mol. Phys., (2016) 1-9. [69] I. Nezbeda, F. Moucka, W.R. Smith, Recent progress in molecular simulation of aqueous electrolytes: force fields, chemical potentials and solubility, Mol. Phys., 114 (2016) 16651690.
AC C
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46
27
ACCEPTED MANUSCRIPT
EP
TE D
M AN U
SC
RI PT
[70] D. Ballal, P. Venkataraman, W.A. Fouad, K.R. Cox, W.G. Chapman, Isolating the non-polar contributions to the intermolecular potential for water-alkane interactions, J. Chem. Phys., 141 (2014) 064905. [71] J.G. McDaniel, A. Yethiraj, Comment on “Isolating the non-polar contributions to the intermolecular potential for water-alkane interactions” [J. Chem. Phys. 141, 064905 (2014)], J. Chem. Phys., 144 (2016) 137101. [72] D. Asthagiri, D. Ballal, P. Venkataraman, W.A. Fouad, K.R. Cox, W.G. Chapman, Response to “Comment on ‘Isolating the non-polar contributions to the intermolecular potential for water-alkane interactions”’ [J. Chem. Phys. 144, 137101 (2016)], J. Chem. Phys., 144 (2016) 137102. [73] Scienomics, MAPS Platform, Paris, France, 2015. [74] J. Ramos, L.D. Peristeras, D.N. Theodorou, Monte Carlo simulation of short chain branched polyolefins in the molten state, Macromolecules, 40 (2007) 9640-9650. [75] A.W. Sousa da Silva, W.F. Vranken, ACPYPE - AnteChamber PYthon Parser interfacE, BMC Res. Notes, 5 (2012) 367-367. [dataset] [76] K.D. Papavasileiou, O.A. Moultos, I.G. Economou, GROMACS IFT input topologies and conformations used in the research article entitled: "Predictions of water/oil interfacial tension at elevated temperatures and pressures: A molecular dynamics simulation study with biomolecular force fields.", Mendeley Data, v2, 2017. http://dx.doi.org/10.17632/y58ffdwkyt.2 [77] E. Lindahl, B. Hess, D. van der Spoel, GROMACS 3.0: a package for molecular simulation and trajectory analysis, J. Mol. Model, 7 (2001) 306-317. [78] D. Van Der Spoel, E. Lindahl, B. Hess, G. Groenhof, A.E. Mark, H.J.C. Berendsen, GROMACS: Fast, flexible, and free, J. Comput. Chem., 26 (2005) 1701-1718. [79] H.J.C. Berendsen, D. van der Spoel, R. van Drunen, GROMACS: A message-passing parallel molecular dynamics implementation, Comput. Phys. Commun., 91 (1995) 43-56. [80] H. Bekker, H.J.C. Berendsen, E.J. Dijkstra, S. Achterop, R. Vondrumen, D. Vanderspoel, A. Sijbers, H. Keegstra, B. Reitsma, M.K.R. Renardus, GROMACS - A Parallel Computer for Molecular-Dynamics Simulations, in: R.A. DeGroot, J. Nadrchal (Eds.) 4th International Conference on Computational Physics (PC 92), World Scientific Publishing, 1993. [81] B. Hess, C. Kutzner, D. van der Spoel, E. Lindahl, GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation, J. Chem. Theory Comput., 4 (2008) 435-447. [82] H.J.C. Berendsen, J.P.M. Postma, W.F. van Gunsteren, A. DiNola, J.R. Haak, Molecular dynamics with coupling to an external bath, J. Chem. Phys., 81 (1984) 3684-3690. [83] T. Darden, D. York, L. Pedersen, Particle mesh Ewald: An N⋅log(N) method for Ewald sums in large systems, J. Chem. Phys., 98 (1993) 10089-10092. [84] R. Salomon-Ferrer, A.W. Götz, D. Poole, S. Le Grand, R.C. Walker, Routine microsecond molecular dynamics simulations with AMBER on GPUs. 2. explicit solvent Particle Mesh Ewald, 9 (2013) 3878-3888. [85] B. Hess, H. Bekker, H.J.C. Berendsen, J.G.E.M. Fraaije, LINCS: A linear constraint solver for molecular simulations, J. Chem. Theory Comput., 18 (1997) 1463-1472. [86] J.S. Rowlinson, B. Widom, Molecular theory of capillarity, Clarendon Press, Oxford, 1982. [87] M. Jorge, M.N.D.S. Cordeiro, Molecular dynamics study of the interface between water and 2-nitrophenyl octyl ether, J. Phys. Chem. B, 112 (2008) 2415-2429.
AC C
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45
28
ACCEPTED MANUSCRIPT
SC
RI PT
[88] S.E. Feller, R.W. Pastor, Constant surface tension simulations of lipid bilayers: The sensitivity of surface areas and compressibilities, J. Chem. Phys., 111 (1999) 1281-1287. [89] W.G. Hoover, Canonical dynamics: Equilibrium phase-space distributions, Phys. Rev. A, 31 (1985) 1695-1697. [90] C.L. Wennberg, T. Murtola, B. Hess, E. Lindahl, Lennard-Jones lattice summation in bilayer simulations has critical effects on surface tension and lipid properties, J. Chem. Theory Comput., 9 (2013) 3527-3537. [91] L. Lundberg, O. Edholm, Dispersion corrections to the surface tension at planar surfaces, J. Chem. Theory Comput., 12 (2016) 4025-4032. [92] N.M. Fischer, P.J. van Maaren, J.C. Ditz, A. Yildirim, D. van der Spoel, Properties of organic liquids when simulated with long-range Lennard-Jones interactions, J. Chem. Theory Comput., 11 (2015) 2938-2944. [93] E.W. Lemmon, M.O. McLinden, D.G. Friend, Thermophysical properties of fluid systems, in: P.J. Linstrom, W.G. Mallard (Eds.), NIST Chemistry WebBook, NIST Standard Reference Database Number 69, National Institute of Standards and Technology, Gaithersburg MD, 20899,10.18434/T4D303, retrieved March 14, 2017.
M AN U
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
AC C
EP
TE D
17
29